Abstract
Despite obvious similarities in their electronic and crystallographic structures, it remains unclear whether the interactions that shape the normal and superconducting (SC) state properties of high-Tc cuprates and infinite-layer nickelates (ILNs) have the same origin. This question has been brought into sharper focus with recent studies on ILNs of improved crystallinity that reveal a SC dome of comparable extent and similar transport properties above Tc as the hole-doped cuprates. The evolution of these properties in the magnetic-field-induced normal state, however, has yet to be determined. Here, we examine the magnetotransport properties of new-generation Nd1−xSrxNiO2 films in the T → 0 limit across the phase diagram in fields up to 54 T. This extensive study reveals that the limiting low-T form of the normal-state resistivity in ILNs exhibits non-Fermi-liquid behaviour over an extended doping range inside the SC dome, rather than at a singular quantum critical point. While there are clear differences in the charge dynamics of ILNs and cuprates, most notably in the magnetoresistance, our findings reveal that both systems exhibit anomalous metallicity characteristic of a quantum critical phase.
Subject terms: Superconducting properties and materials, Electronic properties and materials
Recent advancements in the sample quality of infinite-layer nickelates have made systematic comparisons with cuprates increasingly relevant. Hsu et al. report anomalous electronic transport in high-crystallinity (Nd,Sr)NiO2 at large magnetic fields, map out the transport phase diagram, and discuss similarities and differences with cuprates.
Introduction
In most metals, the resistivity (ρ) follows a power-law temperature (T) or magnetic field (H) dependence in the limiting low-T and low- to intermediate-H regimes. The exponent of this power law is a useful indicator of the nature of the electronic ground state and the dominant interactions that scatter the itinerant electrons. In conventional correlated metals, the low-energy excitations are coherent Landau quasiparticles described within a Fermi-liquid (FL) framework. Accordingly, the change in electrical resistivity Δρ due to varying T and H is both quadratic albeit for different reasons; Δρ(T) and Δρ(H) being proportional to the scattering rate (in this case, inter-electron scattering) and scattering time (and its k-space anisotropy), respectively.
In ‘strange’ metals – epitomised by the high-Tc cuprates – Δρ scales linearly with both T and H over an anomalously broad parameter space, suggesting a breakdown of the quasiparticle description1. An alternative form of anomalous metallicity, often found in correlated metals proximate to a quantum critical point (QCP) of magnetic2 origin, is characterised by a resistivity with an unconventional power-law exponent i.e. Δρ(T) ∝ T n with 1 < n < 2 and, in some cases, a H-linear magnetoresistance (MR)3,4. While the microscopic mechanism responsible for this unconventional behaviour is actively debated, it is an empirical fact that a dome of superconductivity often lies in close proximity to such anomalous metallicity. Examples include the iron-based superconductors5, multilayer graphene superlattices6, and heavy-fermion metals7. It is therefore widely believed that unconventional SC and non-FL behaviours are intrinsically linked. Indeed, the strength of superconductivity is often found to peak near the QCP, motivating the conjecture that critical fluctuations of the electronic order parameter are responsible for both the pairing interaction and the anomalous metallicity7,8.
A SC dome is also realised in the ILNs9–11. Moreover, there are observations of short-ranged spin fluctuations12–15 as well as signatures of charge order16–18 in this system, while interlocking stripe order – known to exist in single-19,20 and tri-layer21,22 rare-earth nickelates – has also been conjectured23, providing an appropriate backdrop for strong correlation physics and anomalous or strange metallicity to emerge. While rapid progress has been made in the improvement of sample quality24,25 and new families have been synthesised10,26–29, information on the limiting form of the normal-state transport properties of ILNs remains sparse, particularly within the temperature and doping range in which superconductivity is observed. This is in large part due to the challenging synthesis conditions required to produce high-quality ILN films, the field strengths required to suppress the superconductivity30,31, and the ubiquitous persistence of an anomalous low-T resistivity upturn32 that complicates the analysis. Very recently, by adopting (LaAlO3)0.3(Sr2TaAlO6)0.7 (LSAT) as the substrate of choice and alleviating the epitaxial strain in the intermediate perovskite phase, a series of highly crystalline Nd1−xSrxNiO2 (NSNO) films largely devoid of extended defects have been prepared over a wide range of Sr doping x24.
This material advancement has motivated a critical examination of the transport properties across the NSNO phase diagram, which now hosts an expanded SC dome (in both T and x) comparable to that found in La2−xSrxCuO4 (LSCO)24. A T-linear in-plane resistivity ρab is also found to persist over an extended T range near the optimal doping level (xopt), with a slope dρab/dT similar to that found near xopt in LSCO24. However, the precise form of ρab(T) beyond xopt and its doping dependence have not yet been ascertained, raising several key questions: does NSNO also exhibit strange metallic behaviour like LSCO (with an extended doping range of non-FL transport properties33) or is the T-linearity seen at higher temperature confined to a singular putative QCP underneath its SC dome? Furthermore, are the low-energy excitations in SC nickelates Landau quasiparticles?
In order to shed more light on these outstanding issues, we have carried out a detailed magnetotransport study on a series of NSNO/LSAT thin films, spanning a broad doping range from slightly underdoped (UD) to highly overdoped (OD), in temperatures down to 0.3 K and magnetic fields up to 54 T. By systematically analysing the high-field magnetotransport data, in combination with previously reported lower-field data24, we uncover the following characteristics of normal-state charge transport in NSNO: 1) the resistivity upturn, while still present, undergoes a qualitative change across xopt, suggesting that the driving force behind the weakly insulating behaviour in UD-NSNO is quenched beyond xopt; 2) the in-plane MR exhibits neither Kohler nor quadrature scaling; rather, the strong T-dependence of the MR appears to reflect the presence of strong k-space anisotropy predominantly in the elastic scattering channel; and 3) the low-T resistivity in the (field-induced) normal state of OD-NSNO (0.175 ≤ x ≤ 0.325) is best described by a non-integer power law Δρ(T) ∝ T n with n ≈ 1.5, consistent with a scenario of dominant magnetic scattering. The T − x phase diagram thus revealed bears signatures found in other unconventional superconductors near a putative QCP34,35, though the persistence of the non-FL form of Δρ(T) over an extended doping range beyond xopt is more reminiscent of the quantum critical phase observed in hole-doped cuprates. The distinct form of Δρ(T) and the MR, however, suggests that the physical processes that drive the anomalous metallicity (and by inference, the superconductivity) in ILNs may be different from those that prevail in their cuprate counterparts.
Results
Transport phase diagram
Figure 1 summarises the evolution of the in-plane resistivity across the NSNO phase diagram. By examining the form of ρab(T) up to 300 K, we deduce a number of key features. Firstly, we find that the narrow window of T-linear resistivity centred around xopt evolves into a broad asymmetric fan of T-linearity bounded from below by an onset temperature T1(x) that tends towards zero as x → xopt. Secondly, the prominent resistive upturn (dρab/dT < 0) found in UD films below a second characteristic temperature T0 is strongly suppressed beyond xopt (Fig. 2). Between T0 and T2 in the UD region, and below T2 in the heavily OD films, Δρab(T) ∝ T 2. In the intermediate regime between xopt and the end of the SC dome, ρab(T) exhibits an anomalous T-dependence that can best be described by a single non-integer power-law with an exponent n = 1.45 ± 0.05 across the entire OD regime. The upper bound of this anomalous form of ρab(T) is also relatively insensitive to x with Tn ~ 50 K. Collectively, these findings reveal that the SC dome of NSNO hosts a state that exhibits anomalous metallicity. Below we examine in detail the evolution of this anomalous metallicity with x, T and H.
Doping endpoint of the correlation-driven resistivity upturn
Previous magnetotransport measurements24 in magnetic fields up to 14 T revealed that the resistive upturn in UD-NSNO is essentially field-insensitive and accompanied by a reduction in the Hall carrier density with decreasing T, suggesting that these upturns are caused by a competing electronic order, such as the charge order recently found in underdoped ILNs16–18. Figure 2a shows the in-plane magnetoresistivity ρab(H) measured between 0.3 and 30 K for samples in the slightly UD to OD regime (0.15 ≤ x ≤ 0.225). The upper critical fields μ0Hc2 inferred using the 50% ρn criterion show a characteristic upturn below 2 K (Fig. 2c), consistent with previous reports31,36. Despite the relatively low Tc, a large magnetic field (>30 T) is required to access the normal state below 1 K (Fig. 2d). By plotting ρab(T) in the field-induced normal state (Fig. 2b) we find a marked reduction in the low-T upturn with no clear doping dependence in its onset temperature T0 or normalised magnitude (<1%). This contrasts with what is observed below xopt24,32 and suggests that the resistivity upturns above and below xopt have different origins. Previous measurements at x ≥ 0.275 found that the resistivity upturn in the highly OD region can be effectively suppressed by magnetic fields and is likely caused, at least in large part, by disorder-induced localisation24. Here, we conclude that the correlation-driven resistivity upturn found in UD-NSNO and the responsible interaction are quenched at xopt.
Magnetoresistance near optimal doping
The large Hc2 values found near xopt require field strengths in excess of 34 T in order to track the normal-state MR far below Tc. Figure 3 shows MR data for three films (x = 0.15, 0.1625, 0.175) obtained in pulsed magnetic fields up to 54 T. An insightful way to distinguish between different forms of MR is to examine the field derivative dρab/dH and its T dependence, as illustrated in the lower panel of Fig. 3a. At T > 35 K, dρab/dH ∝ H, confirming that the MR follows a conventional H2 behaviour. Below 35 K, however, dρab/dH develops a marked curvature, most clearly revealed in the trace at T = 25 K. Despite the apparent downturn in dρab/dH with increasing H, at no point does dρab/dH become constant, indicating that the strictly H-linear MR ubiquitously found across the strange metal regime in hole-doped cuprates37 is not realised in ILNs, at least in the field range studied.
Another notable feature of Fig. 3a is the marked increase in dρab/dH with decreasing T. Indeed, the slope of the MR (i.e. magnitude of dρab/dH) increases by more than one order of magnitude between 95 and 4.2 K, despite the fact that the zero-field resistivity ρ(0) changes by only a factor of 2, leading to a strong violation of the Kohler scaling observed in conventional metals (Supplementary Fig. 4). Within a Boltzmann theoretical framework, the rapid increase in MR magnitude with decreasing T can only be reconciled by invoking a corresponding and marked increase in the k-space anisotropy of the mean-free-path ℓ, predominantly in the elastic scattering channel (see further details in Supplementary Note C). Alternative functional forms to describe the normal-state MR behaviour, including a standard two-carrier model (Δρ(H) = AH 2/(1 + BH 2)) or a dual-component model (Δρ(H) = β1H + β2H 2), fail to capture the MR data over an appreciable T and H window (Supplementary Fig. 5). The absence of a convergence of dρab/dH curves at the high-field limit (Fig. 3a and Supplementary Fig. 10) also indicates that the limiting low-T MR fails to follow the quadrature form observed in hole-doped cuprates37 and quantum critical metals3,4 (Supplementary Fig. 4).
Empirically, we find the MR data (at least above 10 T) can be best described by a single power law Δρ(H) = βmH m for all three x ≈ xopt. As demonstrated in Fig. 3b via log-log plots of the fractional MR Δρ(H)/ρ(0), m evolves smoothly from 2 at T≥ 50 K to ≈ 1.5 at 25 K, above which ρ(0) is experimentally known. Applying the power-law MR analysis to the field-induced normal state below Tc (as inferred from the field scale above which ρ(H) shows a positive curvature; Supplementary Figs. 6–9), we find that m converges to an anomalous value 1.37 ± 0.06 for all three dopings near xopt (Fig. 3d), suggesting an anomalous nature of the metallic ground state at x ≈ xopt that is distinct from that of the cuprates or other quantum critical metals.
Non-Fermi liquid zero-field resistivity
The diminished magnitude (≲1%) and temperature scale (≲4 K) of the resistive upturn for x ≥ xopt provides an opportunity to investigate the evolution of ρab(T) in NSNO over a more extended temperature and doping range than previous studies24. Figure 4 shows ρab(T) measured at zero and fixed field strengths (μ0H = 34 T for 0.175 ≤ x ≤ 0.25; 14 T for x = 0.275 (from previous work24); 12 T for x ≥ 0.30). The x = 0.275 film has the lowest T0 = 0.96 K with a vanishingly small upturn and a negligible MR at 14 T, providing the best opportunity to examine the functional form of normal-state ρ(T) as T → 0. Note that the 0-T and 14-T data are in good agreement within the common temperature range, indicating that the application of magnetic field reveals the normal-state behaviour of ρab(T) below Tc without affecting its functional form.
Inspection of dρab/dT (Fig. 4a) reveals a small kink at 3 K, below which dρab/dT is T-linear (within noise level) with a near zero 0-K intercept. This is consistent with Δρ(T) ∝ T 2 albeit over a narrow T-window. An alternative functional form incorporating both T-linear and T 2 components: ρ(T) = ρ0 + α1T + α2T 2, used previously to describe the limiting low-T resistivity of hole-doped cuprates33, is found to hold over the T-window 3 K ≤T≤ 18 K with a corresponding T-linear behaviour in dρab/dT and a finite 0-K intercept. Indeed, the ratio of the low- and high-T T-linear coefficients in NSNO and LSCO are found to show a very similar evolution with x within experimental uncertainty (Supplementary Note A.2 and Supplementary Fig. 3). For the T-linear term to be intrinsic, the kink in dρab/dT below 3 K would have to reflect the beginning of influence of the resistive upturn. On the other hand, if the pronounced downward curvature in dρab/dT is intrinsic, then ρab(T) can be better described by a power-law function: ρ(T) = ρ0 + αnT n with 1 < n < 2 over a wider T-window. To extract the power-law exponent n with any possible impact of the residual upturn minimised, we fit dρab/dT(μ0H = 14 T) using dρab/dT = nαnT n−1 within a T-window between 3T0 and 18 K (the highest temperature measured). The extracted n = 1.41 ± 0.02 from the 14-T data is found to describe dρab/dT(μ0H = 0) up to 60 K. We note that varying the lower T-limit for the power-law fitting between 2T0 and 5T0 does not affect the extracted n, as nαnT n−1 is constrained to go through the origin, demonstrating that the extraction of the resistivity power-law parameters is robust despite the presence of these small resistive upturns in OD-NSNO.
By examining ρab(T) of OD films with 0.25 ≥ x ≥ 0.175, we again find for all x that the dρab/dT curves measured at 34 T exhibit a pronounced downward curvature and connect smoothly to the 0-T data in the normal state (Fig. 4b). Overall, the power-law function can describe the experimental data over the widest temperature range among the three functional forms considered (Supplementary Figs. 13 and 14) and, remarkably, the extracted n for 0.175 ≤ x ≤ 0.325 converges within a narrow range n = 1.45 ± 0.05 (Fig. 4c). The approximate linear-in-T 1.45 behaviour of the normal-state ρab(T) can also be seen in Fig. 4e, where the residual-free resistivity Δρab(T) is plotted against T 1.45 for all OD films. Interestingly, the power-law resistivity prefactor αn is found to correlate to the residual resistivity ρ0 more strongly than to x (Supplementary Fig. 14). Thus, we conclude that the low-T ρab(T) in OD-NSNO is best described by a power-law behaviour ρ(T) = ρ0 + αnT n, with an unconventional n ≈ 1.5, though the alternative T + T 2 form reminiscent of the hole-doped cuprates cannot be definitely ruled out. Irrespective of this, the extended doping range over which the low-T ρab(T) exhibits non-FL behaviour is inconsistent with scenarios based on a standard QCP and is more reflective of the quantum critical phase and strange metal physics manifest in the cuprates.
Discussion
Through a systematic investigation over a wide parameter space, we have uncovered the anomalous resistivity and MR behaviour of NSNO and mapped out its evolution across the phase diagram. While we are not aware of existing theoretical models that could produce the anomalous MR power law found in optimally doped NSNO at T ≲ Tc, similar behaviour has previously been observed in other superconductors38,39. To investigate whether such a MR behaviour can be explained within a conventional theoretical (Boltzmann) framework, we performed a series of numerical simulations using calculated Fermi surface parameterisation for the superconducting nickelates40 and various models to describe the k-space scattering rate anisotropy:
1 |
where 1/τ is the inverse of the carrier lifetime (i.e. scattering rate), ϕ is the angle around the Fermi surface, 1/τ0 is the isotropic (k-independent, inelastic) scattering rate, and 1/τk is the anisotropic (k-dependent, elastic) scattering rate. We find that the normal-state MR with an apparent power-law behaviour can be qualitatively reproduced using 1/τ(ϕ) models with a large scattering rate anisotropy (Supplementary Fig. 11 and Supplementary Table 1). The significant increase in the MR magnitude with decreasing T (Fig. 3) thus reflects a rapid increase in the scattering rate anisotropy and, given the modest change in zero-field resistivity (ρ(95 K)/ρ(4.2 K) ≲ 2), a dramatic increase in / exceeding one order of magnitude. Moreover, we find that the best quantitative agreement is achieved by excluding the contribution from the putative rare-earth (RE) Fermi pocket, implying that the RE band contributes minimally to the charge transport, at least for x ≈ xopt. While the power-law MR can be (at least) qualitatively captured using the Boltzmann framework, we note that these simulations only serve to illustrate one possible scenario for the observed MR in NSNO, and the precise Fermi surface geometry and scattering rate anisotropy remain to be established.
A T 1.5 resistivity power law has been observed in a variety of correlated electron systems with magnetic order, including heavy-fermion superconductors41–43, d-electron itinerant ferromagnets (IFMs)44–46, and metallic spin glasses (SGs)47,48. In heavy-fermion systems, a T 1.5 resistivity is typically found within a fan-like region centred around a putative antiferromagnetic QCP, sandwiched by regions of T 2 resistivity. In contrast, in IFM and SG, T 1.5 resistivity is found as T → 0 over an extended region of pressure44,45 or magnetic dopant47. While at present we cannot rule out the presence of an antiferromagnetic QCP at x ≈ 0.275, we find no enhancement in the prefactor of T 2 resistivity as x → 0.275 (assuming ρ(T) = ρ0 + A2T 2 for the limiting low-T behaviour; Supplementary Fig. 14) which would signify a divergence of the quasiparticle effective mass upon approach to a putative QCP2,34. (We note here that the low-T resistivity plateau found in the T 2 resistivity regime in x≥0.30, with a small negative MR, is unlikely to be associated with short-lived superconducting puddles as found in other superconductors in the 2D limit49,50, for which a large positive MR is expected.) The IFMs known to exhibit a T 1.5 resistivity (beyond the pressure-tuned ferromagnetic QCP), namely MnSi and ZrZn2, lie well within the clean limit with a residual resistivity ρ0 ≲ 0.5 μΩ cm (refs. 44,45), two orders of magnitude lower than that of the current NSNO series. Moreover, the power-law exponent of IFMs is highly sensitive to applied magnetic field and approaches 2 at μ0H > 1 T (refs. 51,52). These distinct differences suggest that the interactions responsible for the T 1.5 resistivity in IFMs are unlikely to be manifest in NSNO. Metallic SGs, on the other hand, are disordered systems (ρ0 ≫ 10 μΩ cm) without long-range magnetic order. Notably, in some SGs, a T 2 resistivity is recovered at high doping levels as T → 0 (ref. 47), as found here in heavily OD-NSNO. We also find approximate T 1.5 resistivity in overdoped Pr1−xSrxNiO2 (x = 0.24) and La1−xSrxNiO2 (x = 0.28; Supplementary Fig. 16), implying that the unconventional resistivity power law is intrinsic to the Ni-O lattice and not influenced by the respective lanthanide ion. Given that signatures of short-range spin order have been observed in ILNs13–15, we postulate that the observation of an extended non-FL power-law resistivity in superconducting nickelates could be a consequence of their metallic SG state.
Typically, a low-T resistivity power law with an unconventional exponent (n < 2) is considered to be manifestation of a FL instability and the associated low-energy excitations are no longer coherent quasiparticles; the electrical current in these unconventional metals originates from diffusive motion of the interacting electrons2,53. In the case of a metallic SG, however, it has been theoretically shown that there the quasiparticles remain intact despite the appearance of non-FL behaviour in the electrical transport54,55. The specific heat coefficient and T 2 resistivity in the disordered state in turn are not critically enhanced. Nonetheless, the unconventional power law supports the strongly correlated nature of the itinerant electrons and hints at the coexistence of SG and SC in the ILNs15. Importantly, the manifestation of approximate T 1.5 resistivity power law provides evidence for dominant magnetic scatterings in the normal ground state, and supports the existence of a fluctuating magnetic order intrinsic to the Ni-O lattice12,15 under hole doping.
It has been suggested that the presence of RE Fermi pockets leads to an effective hole doping into the Ni-O lattice56 that is responsible for the absence of static antiferromagnetic order in the nominally undoped ILNs. If true, the similar dρab/dT observed in NSNO and LSCO near their respective xopt are then coincidental, and the comparison of the doping effect should be made between LSCO at x ≈ 0.15 and NSNO x ≳ 0.20. (The fact that the present NSNO thin films exhibit nearly identical dρab/dT values as found in high-quality LSCO single crystals strongly suggests that neither extended defects – known to affect the transport properties of earlier NSNO films – and/or dominant electron-phonon scattering are playing a major role in quantifying the in-plane resistivity values of these next-generation films; see Supplementary Notes A and F for further discussions.) Nevertheless, in all OD-NSNO investigated in this work, we do not observe a predominantly T-linear resistivity (over an extended T-window at low T; Supplementary Fig. 14), suggesting the potential presence of an additional band is not the leading cause for the absent strange metallicity. Secondly, if the low-energy excitations in the low-T normal state of superconducting ILNs are indeed quasiparticles, as is the case in SGs, it may explain certain aspects of the transport characteristics of OD-NSNO. In this scenario, xopt would not correspond to a conventional QCP at which a T-linear resistivity as T → 0 is found and the quasiparticles are destroyed. Nonetheless, experimental signatures of a low-T SG state have also been detected in LSCO near optimal doping57, hinting that a SG state and strange metallicity are not mutually exclusive. Finally, a series of recent experiments15–18,58 has found that the landscape of competing orders in the ILNs is considerably different than that in the cuprates59. It is therefore likely that an intricate interplay between the various electronic orders (further discussions in Supplementary Note F) play a dominant role in determining the transport characteristics (and mediating superconductivity) in the ILNs.
While the exact mechanism responsible for the anomalous metallicity remains to be identified, our results highlight the impact of magnetic interactions in governing the charge transport of normal-state electrons. We thus expect our present study lays important groundwork for understanding of quantum phenomena in the ILNs and provides useful constraints for a viable theoretical model for the anomalous metallicity (and superconductivity) emerging upon doping the NiO2 planar square lattice.
Methods
Thin film growth
NSNO thin films of approximately 15 unit cells were grown on LSAT (001) substrates with a SrTiO3 capping layer of approximately 4 unit cells using pulsed laser deposition technique24. We confirm that, other than the additionally synthesised x = 0.1625 sample, the thin film samples presented here are the exact same samples that were studied in Lee et al.24.
Magnetotransport measurements
Gold wires of 25-μm diameter were attached to ultrasonically bonded Al leads using DuPont 4929N silver paint. An AC current of 5 or 10 μA with a frequency of 13–30 Hz was used in DC field measurements and 5–10 kHz for pulsed field measurements. Magnetic fields, applied parallel to the crystalline c-axis, were generated up to 14 T using a Physical Properties Measurement System by Quantum Design Inc., up to 34 T using a Florida-Bitter magnet coupled to a 3He refrigerator at the High Field Magnet Laboratory in Radboud University, the Netherlands, and up to 54 T using a pulsed-field magnet at the Laboratoire National des Champs Magnétiques Intenses in Toulouse, France.
Tight-binding model
We adopt the tight-binding model developed in a recent study on the electronic structure of superconducting five-layer nickelate Nd6Ni5O1240. The Fermi surface of Nd6Ni5O12 contains five sheets derived from the Ni band and one sheet from the Nd d band. The Ni-3 and Ni-1 bands (following the notation of Table I in ref. 40) are chosen to simulate the Fermi surface of Nd1−xSrxNiO2 near optimal doping, given the current lack of experimentally established doping evolution of Fermi surface in this system, based on their close similarity to the DFT-derived Fermi surface of LaNiO2 at kz = 0 and kz = π/c, respectively60.
Magnetoresistivity simulations
We calculate the magnetoresistivity using the Chambers solution to the Boltzmann transport equation with generalised to quasi-2D systems with anisotropic scatterings:
2 |
where ∫ d 3k is an integral over the first Brillouin zone, D(k) is the density of states at point k, vi is the ith component of the quasiparticle velocity, and is an integral over the quasiparticle lifetime τ. The scattering rate anisotropy enters through 1/τ(ϕ) parametrised by:
3 |
where ϕ is the azimuthal angle around the Fermi surface, 1/τ0 is the ϕ-independent isotropic scattering rate, 1/τk is the maximum anisotropic scattering rate found in the Brillouin zone (i.e. ’hot spots’), f is the function that characterises the scattering rate anisotropy (i.e. sine/cosine function for the nodal/antinodal hotspot model, respectively), and ν is a variable exponent (ν = 4, 8, 10, or 12). Note that no scattering rate anisotropy is applied to the Nd pocket. For the MR simulations including both the kz = 0 and kz = π/c Ni sheets, the contribution to the total conductivity is halved to maintain the total carrier number close to half-filling; for MR simulations including the Nd pocket, the entire contribution from the Nd pocket is added into the total conductivity.
Supplementary information
Acknowledgements
We thank G. Grissonnanche and C.-H. Chung for stimulating discussions, and D. Vignolles and C. Proust for experimental support at LNCMI. We acknowledge the support of HFML-RU/NWO-I and LNCMI-T, both members of the European Magnetic Field Laboratory (EMFL). This work was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (Grant agreement no. 835279-Catch-22) (Y.-T.H., C.D., N.E.H.). This work was also supported by the US Department of Energy, Office of Science, Basic Energy Sciences, Materials Sciences and Engineering Division under contract no. DE-AC02-76SF00515 (synthesis and transport measurements) and the Gordon and Betty Moore Foundation’s Emergent Phenomena in Quantum Systems Initiative (Grant no. GBMF9072, synthesis equipment) (K.L., B.Y.W., H.Y.H.). This work is partially supported by the Yushan Fellow Program by the Ministriy of Education (MOE), Taiwan (MOE-112-YSFMS-0002-002-P1) (Y.-T.H.) and a Radboud Excellence Fellowship by Radboud University (S.B.).
Author contributions
Y.-T.H., K.L., H.Y.H. and N.E.H. initiated the project. K.L. and B.Y.W. synthesised and characterised the nickelate thin films. Y.-T.H., S.B., C.D., A.C., A.K., and I.H-D. performed the high-field magnetotransport measurements. Y.-T.H., C.D., A.C., and N.E.H. analysed the data. Y.-T.H. and N.E.H. wrote the manuscript with input from all authors.
Peer review
Peer review information
Nature Communications thanks the anonymous reviewers for their contribution to the peer review of this work. A peer review file is available.
Data availability
The data that support the findings of this study are available from the Dryad data repository (10.5061/dryad.8sf7m0czw). Any additional data are available from the corresponding authors upon request.
Code availability
The codes associated with the band structure and transport coefficient calculations that support this study are available from the corresponding authors upon request.
Competing interests
The authors declare no competing interests.
Footnotes
Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Contributor Information
Yu-Te Hsu, Email: ythsu@phys.nthu.edu.tw.
Nigel E. Hussey, Email: n.e.hussey@bristol.ac.uk
Supplementary information
The online version contains supplementary material available at 10.1038/s41467-024-54135-y.
References
- 1.Phillips, P. W., Hussey, N. E. & Abbamonte, P. Stranger than metals. Science377, 169 (2022). [DOI] [PubMed] [Google Scholar]
- 2.Löhneysen, vH., Rosch, A., Vojta, M. & Wölfle, P. Fermi-liquid instabilities at magnetic quantum phase transitions. Rev. Mod. Phys.79, 1015–1075 (2007). [Google Scholar]
- 3.Hayes, I. M. et al. Scaling between magnetic field and temperature in the high-temperature superconductor BaFe2(As1−xPx)2. Nat. Phys.12, 916–919 (2016). [Google Scholar]
- 4.Licciardello, S. et al. Coexistence of orbital and quantum critical magnetoresistance in FeSe1−xSrx. Phys. Rev. Res.1, 023011 (2019). [Google Scholar]
- 5.Shibauchi, T., Carrington, A. & Matsuda, Y. A quantum critical point lying beneath the superconducting dome in iron pnictides. Annu. Rev. Condens. Matter Phys.5, 113–135 (2014). [Google Scholar]
- 6.Cao, Y. et al. Strange metal in magic-angle graphene with near Planckian dissipation. Phys. Rev. Lett.124, 076801 (2020). [DOI] [PubMed] [Google Scholar]
- 7.Gegenwart, P., Si, Q. & Steglich, F. Quantum criticality in heavy-fermion metals. Nat. Phys.4, 186–197 (2008). [Google Scholar]
- 8.Sachdev, S. & Keimer, B. Quantum criticality. Phys. Today64, 29–35 (2011). [Google Scholar]
- 9.Li, D. et al. Superconductivity in an infinite-layer nickelate. Nature572, 624 (2019). [DOI] [PubMed] [Google Scholar]
- 10.Li, D. et al. Superconducting dome in Nd1−xSrxNiO2 infinite layer films. Phys. Rev. Lett.125, 027001 (2020). [DOI] [PubMed] [Google Scholar]
- 11.Zeng, S. et al. Phase diagram and superconducting dome of infinite-layer Nd1−xSrxNiO2 thin films. Phys. Rev. Lett.125, 147003 (2020). [DOI] [PubMed] [Google Scholar]
- 12.Lu, H. et al. Magnetic excitations in infinite-layer nickelates. Science373, 213–216 (2021). [DOI] [PubMed] [Google Scholar]
- 13.Cui, Y. et al. NMR evidence of antiferromagnetic spin fluctuations in Nd0.85Sr0.15NiO2. Chin. Phys. Lett.38, 067401 (2021). [Google Scholar]
- 14.Lin, H. et al. Universal spin-glass behavior in bulk LaNiO2, PrNiO2, and NdNiO2. N. J. Phys.24, 013022 (2022). [Google Scholar]
- 15.Fowlie, J. et al. Intrinsic magnetism in superconducting infinite-layer nickelates. Nat. Phys.18, 1043–1047 (2022). [Google Scholar]
- 16.Krieger, G. et al. Charge and spin order dichotomy in NdNiO2 driven by the capping layer. Phys. Rev. Lett.129, 027002 (2022). [DOI] [PubMed] [Google Scholar]
- 17.Rossi, M. et al. A broken translational symmetry state in an infinite-layer nickelate. Nat. Phys.18, 869–873 (2022). [Google Scholar]
- 18.Tam, C. C. et al. Charge density waves in infinite-layer NdNiO2 nickelates. Nat. Mater.21, 1116–1120 (2022). [DOI] [PubMed] [Google Scholar]
- 19.Sugai, S., Sato, M., Kobayashi, M., Akimitsu, J. & Takagi, H. High-energy spin excitations in the insulating phases of high-Tc superconducting cuprates and La2NiO4. Phys. Rev. B42, 1045 (1990). [DOI] [PubMed] [Google Scholar]
- 20.Sachan, V., Buttrey, D. J., Tranquada, J. M., Lorenzo, J. E. & Shirane, G. Charge and spin ordering in La2−xSrxNiO4.00 with x = 0.135 and 0.20. Phys. Rev. B51, 12742 (1995). [DOI] [PubMed]
- 21.Zhang, J. et al. Stacked charge stripes in the quasi-2D trilayer nickelate La4Ni3O8. Proc. Natl Acad. Sci.113, 8945–8950 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Zhang, J. et al. Spin stripe order in a quare planar trilayer nickelate. Phys. Rev. Lett.122, 247201 (2019). [DOI] [PubMed] [Google Scholar]
- 23.Zhang, R. et al. Peierls distortion driven multi-orbital origin of charge density waves in the undoped infinite-layer nickelate. Preprint at https://arxiv.org/abs/2207.00184.
- 24.Lee, K. et al. Linear-in-temperature resistivity for optimally superconducting (Nd, Sr)NiO2. Nature619, 288–292 (2023). [DOI] [PubMed] [Google Scholar]
- 25.Osada, M., Fujiwara, K., Nojima, T. & Tsukazaki, A. Improvement of superconducting properties in La1−xSrxNiO2 thin films by tuning topochemical reduction temperature. Phys. Rev. Mater.7, L051801 (2023). [Google Scholar]
- 26.Osada, M. et al. A superconducting praseodymium nickelate with infinite layer structure. Nano Lett.20, 5735 (2020). [DOI] [PubMed] [Google Scholar]
- 27.Osada, M. et al. Nickelate superconductivity without rare-earth magnetism: (La, Sr)NiO2. Adv. Mater.33, 2104083 (2021). [DOI] [PubMed]
- 28.Zeng, S. W. et al. Superconductivity in infinite-layer nickelate La1−xCaxNiO2 thin films. Sci. Adv.8, eabl9927 (2022). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Wei, W., Vu, D., Zhang, Z., Walker, F. J. & Ahn, C. H. Superconducting Nd1−xEuxNiO2. Sci. Adv.9, eadh3327 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Chow, L. E. et al. Pauli-limit violation in lanthanide infinite-layer nickelate superconductors. Preprint at https://arxiv.org/abs/2204.12606.
- 31.Wang, B. Y. et al. Effects of rare-earth magnetism on the superconducting upper critical field in infinite-layer nickelates. Sci. Adv.9, eadf6655 (2023). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 32.Hsu, Y.-T. et al. Insulator-to-metal crossover near the edge of the superconducting dome in Nd1−xSrxNiO2. Phys. Rev. Res.3, L042015 (2021). [Google Scholar]
- 33.Cooper, R. A. et al. Anomalous criticality in the electrical resistivity of La2−xSrxCuO4. Science323, 603–607 (2009). [DOI] [PubMed] [Google Scholar]
- 34.Analytis, J. G. et al. Transport near a quantum critical point in BaFe2(As1−xPx)2. Nat. Phys.10, 194–197 (2014). [Google Scholar]
- 35.Licciardello, S. et al. Electrical resistivity across a nematic quantum critical point. Nature567, 213–217 (2019). [DOI] [PubMed] [Google Scholar]
- 36.Wang, B. Y. et al. Isotropic Pauli-limited superconductivity in the infinite-layer nickelate Nd0.775Sr0.225NiO2. Nat. Phys.17, 473–477 (2021). [Google Scholar]
- 37.Ayres, J. et al. Incoherent transport across the strange-metal regime of overdoped cuprates. Nature595, 661–666 (2021). [DOI] [PubMed] [Google Scholar]
- 38.Chen, B. et al. Large magnetoresistance and superconductivity in α-gallium single crystals. npj Quantum Mater.3, 40 (2018). [Google Scholar]
- 39.Poniatowski, N., Sakar, T. & Greene, R. L. Anomalous normal-state magnetotransport in an electron-doped cuprate. Phys. Rev. B103, 125102 (2021). [Google Scholar]
- 40.Grissonnanche, G. et al. Seebeck coefficient in a nickelate superconductor: electronic dispersion in the strange metal phase. Preprint at https://arxiv.org/abs/2210.10987.
- 41.Stewart, G. R. Addendum: Non-Fermi-liquid behavior in d- and f-electron metals. Rev. Mod. Phys.78, 743–753 (2006). [Google Scholar]
- 42.Zou, Y. et al. Fermi liquid breakdown and evidence for superconductivity in YFe2Ge2. Phys. Satus Solidi RRL8, 928–930 (2014). [Google Scholar]
- 43.Tomita, T., Kuga, K., Uwatoko, Y., Coleman, P. & Nakatsuji, S. Strange metal without magnetic criticality. Science349, 506–509 (2015). [DOI] [PubMed] [Google Scholar]
- 44.Doiron-Leyraud, N. et al. Fermi-liquid breakdown in the paramagnetic phase of a pure metal. Nature425, 595–599 (2003). [DOI] [PubMed] [Google Scholar]
- 45.Smith, R. P. et al. Marginal breakdown of the Fermi-liquid state on the border of metallic ferromagnetism. Nature455, 1220–1223 (2008). [Google Scholar]
- 46.Brando, M. et al. Logarithmic Fermi-liquid breakdown in NbFe2. Phys. Rev. Lett.101, 026401 (2008). [DOI] [PubMed] [Google Scholar]
- 47.Ford, P. J. & Mydosh, J. A. Electrical resistivity of noble-metal-host–3d solute spin-glass alloys. Phys. Rev. B14, 2057–2069 (1976). [Google Scholar]
- 48.Miranda, E. & Dobrosavljevic, V. Disorder-driven non-Fermi liquid behaviour of correlated electrons. Rep. Prog. Phys.68, 2337–2408 (2005). [Google Scholar]
- 49.Kapitulnik, A., Kivelson, S. A. & Spivak, B. Colloquium: Anoamlous metals: Failed superconductors. Rev. Mod. Phys.91, 011002 (2019). [Google Scholar]
- 50.Yang, C. et al. Intermediate bosonic metallic state in the superconductor-insulator transition. Science366, 1505–1509 (2019). [DOI] [PubMed] [Google Scholar]
- 51.Kabeya, N. et al. Non-Fermi liquid state bounded by a possible electronic topological transition in ZrZn2. J. Phys. Soc. Jpn.81, 073706 (2012). [Google Scholar]
- 52.Ritz, R. et al. Formation of a topological non-Fermi liquid in MnSi. Nature497, 231 (2013). [DOI] [PubMed] [Google Scholar]
- 53.Hartnoll, S. A. & Mackenzie, A. P. Colloquium: Planckian dissipation in metals. Rev. Mod. Phys.94, 041002 (2022). [Google Scholar]
- 54.Sengupta, A. M. & Georges, A. Non-Fermi-liquid behavior near a T = 0spin-glass transition. Phys. Rev. B52, 10295 (1995). [DOI] [PubMed] [Google Scholar]
- 55.Sachdev, S. & Read, N. Metallic spin glasses. J. Phys. Condens. Matter8, 9723 (1996). [DOI] [PubMed] [Google Scholar]
- 56.Zhang, G.-M., Yang, Y.-F. & Zhang, F.-C. Self-doped Mott insulator for parent compound of nickelate superconductors. Phys. Rev. B101, 020501(R) (2020). [Google Scholar]
- 57.Frachet, M. et al. Hidden magnetism at the pseudogap critical point of a cuprate supercondudctor. Nat. Phys.16, 1064–1068 (2020). [Google Scholar]
- 58.Poniatowski, N., Sakar, T. & Greene, R. L. Absence of 3a0 charge density wave order in the infinite-layer nickelate NdNiO2. Nat. Mater.103, 125102 (2021). [Google Scholar]
- 59.Benckiser, E., Hepting, M. & Keimer, B. Neighbours in charge. Nat. Mater.21, 1102–1103 (2022). [DOI] [PubMed] [Google Scholar]
- 60.Botana, A. S. & Norman, M. R. Similarities and differences between LaNiO2 and CaCuO2 and implications for superconductivity. Phys. Rev. X10, 011024 (2020). [Google Scholar]
- 61.Hussey, N. E. et al. Dichotomy in the T-linear resistivity in hole-doped cuprates. Philos. Trans. R. Soc. A369, 1626–1639 (2011). [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
The data that support the findings of this study are available from the Dryad data repository (10.5061/dryad.8sf7m0czw). Any additional data are available from the corresponding authors upon request.
The codes associated with the band structure and transport coefficient calculations that support this study are available from the corresponding authors upon request.